Вы находитесь на странице: 1из 11

Entropy 2013, 15, 1221-1231; doi:10.

3390/e15041221
OPEN ACCESS

entropy
ISSN 1099-4300
www.mdpi.com/journal/entropy
Article

Derivation of 2D Power-Law Velocity Distribution Using


Entropy Theory
Vijay P. Singh 1,2,*, Gustavo Marini 3 and Nicola Fontana 3
1

Department of Biological and Agricultural Engineering, Texas A and M University, 2117 TAMU,
College Station, TX 77842, USA
Department of Civil and Environmental Engineering, Texas A and M University, 2117 TAMU,
College Station, TX 77842, USA
Dipartimento di Ingegneria, Universit degli Studi del Sannio, Piazza Roma 21, 82100, Benevento,
Italy; E-Mails: gustavo.marini@unisannio.it (G.M.); fontana@unisannio.it (N.F.)

* Author to whom correspondence should be addressed; E-Mail: vsingh@tamu.edu.


Received: 1 February 2013; in revised form: 1 April 2013 / Accepted: 2 April 2013 /
Published: 8 April 2013

Abstract: The one-dimensional (1D) power law velocity distribution, commonly used for
computing velocities in open channel flow, has been derived empirically. However, a
multitude of problems, such as scour around bridge piers, cutoffs and diversions, pollutant
dispersion, and so on, require the velocity distribution in two dimensions. This paper
employs the Shannon entropy theory for deriving the power law velocity distribution in
two-dimensions (2D). The development encompasses the rectangular domain, but can be
extended to any arbitrary domain, including a trapezoidal domain. The derived
methodology requires only a few parameters and the good agreement is confirmed by
comparing the velocity values calculated using the proposed methodology with values
derived from both the 1D power law model and a logarithmic velocity distribution
available in the literature.
Keywords: entropy; flow measurement; open-channel flow; Shannon entropy; streamflow;
velocity distribution

Entropy 2013, 15

1222

1. Introduction
Fundamental to hydraulic modeling of natural rivers, including modeling of sediment and
contaminant transport, design of channels and river training works, design of hydraulic structures,
development of a rating curve, and so on, is the velocity distribution. In general, velocity in open
channels varies in three dimensions: along flow depth, width, and length. In a given cross-section, the
velocity varies along the flow depth and along the transverse direction or width. Along the flow depth
it varies from zero at the channel bed to a maximum value which, depending on channel geometric
characteristics, exists either at the water surface or some distance below it. Along the transverse
direction, it varies from zero at one boundary to a maximum value at some point in the flow domain
and then declines to zero at the other boundary. The non-uniformity in the distribution of velocity from
the bed to the open water surface and from one boundary to the other is caused by shear stresses.
It is known that velocity is subject to the uncertainties due to both the intrinsic randomness and our
inability to interpret the complex interactions. The empirical 1D power law velocity distribution is not
capable of incorporating these uncertainties. More modern approaches, mainly based on the concept of
entropy, as in the fields of hydraulics and hydrology [1,2], consider velocity as a probabilistic variable,
taking into account this uncertainty. Following Chiu [3], it is plausible to consider time-averaged
velocity as a probabilistic variable, derive the probability distribution of velocity and then derive the
velocity distribution. Chiu [3] proposed the entropy theory for deriving the velocity distribution which
has since been employed by Chiu and his associates [310], as well others [1117]. Some of these
investigations dealt with 1D and some with 2D velocity distributions. Although the 2D velocity
distribution proposed by Chiu [4] has been used in a number of theoretical investigations, its practical
usefulness is inhibited by the many parameters it contains. Moramarco et al. [18] simplified the 2D
model by applying it to several verticals, but still it contains too many parameters. Also employing the
entropy theory, Marini et al. [19] developed a generic 2D velocity distribution, which involves the
geometry of the cross-section, average velocity, and the position and value of maximum velocity. The
model does not require calibration and was successfully tested by comparing theoretical values with
experimental measurements [20].
This approach is also employed in the present paper, aiming at deriving the 2D power law velocity
distribution using the Shannon entropy and comparing it with the entropy-based logarithmic velocity
distribution developed by Marini et al. [19].
2. Derivation of 2D Power Law Velocity Distribution
It is assumed that the time-averaged velocity can be considered as a random variable [21]. The
entropy-based approach for deriving the 2D power law velocity distribution entails maximizing the
Shannon entropy subject to specified constraints and hypothesizing a relation between the cumulative
distribution function (CDF) of u and flow depth y. The Shannon entropy [22] of velocity u, S(u), can
be expressed as:
S (u ) S [ f ( u )]

u max

f (u ) ln
0

f ( u ) du

(1)

Entropy 2013, 15

1223

where f(u) is the probability density function (PDF) of u, and umax is the maximum velocity. Two
constraints must been defined to derive the 1D power law velocity distribution:
u max

f (u )du 1

(2)

ln(u ) f (u )du ln u

(3)

which defines the PDF of u [5], and:

u max

which represents the mean of the logarithmic velocity [23]. Since the constraints given by Equations
(2) and (3) also hold in the 2D domain, a general equation for the velocity PDF was obtained.
Following the approach proposed by Singh [24] for 1D domain, maximizing entropy according to the
principle of maximum entropy (POME) [2527] and using the method of Lagrange multipliers, the
entropy-based PDF of u can be written as:
(4)
f (u ) exp u 1
1

where 1 and 2 are the Lagrange multipliers, calculated according to Equations (2) and (3).
To derive the 2D velocity distribution, let us consider a 2D domain (x, y), with x denoting the
transverse direction and y the vertical direction measured from the bed, upward positive. Let u = u(x, y)
be the velocity distribution, fu(x, y) the PDF and Fu(x, y) the CDF. It is convenient to assume
v = ln(u) [i.e., u = exp(v)]. Following the methodology developed by Marini et al. [19] and taking the
partial derivatives of F(u) with respect to x and y, we obtain:

v f e e
x
v f e e
y

F u F e v
dF e v d e v

x
dv
x
de v
v
v
dF e d e v
F u F e

dv
y
y
de v

v
x
v
y

(5)

Using Equation (4), Equations (5) can be rewritten as:


F u
exp 1 2 v 1e v
x
F u
exp 1 2 v 1e v
x

v
v
exp 1 2 1v 1
x
x
v
v
exp 1 2 1v 1
y
y

Denoting 2+1 = n, Equations (6) can be rewritten as:


F (u )
v
exp 1 1 exp nv
x
x
F (u )
v
exp 1 1 exp nv
x
y
Setting exp(nv) = w, the partial derivatives of w with respect to x and y can be cast as:
w
v
n exp nv
x
x
w
v
n exp nv
y
y
Substituting Equations (8) into Equations (7), the following system of equations is obtained:

(6)

(7)

(8)

Entropy 2013, 15

1224

F (u )
w
n exp1 1
x
x
F (u )
w
n exp1 1
y
y

(9)

Equations (9) can be integrated using the Leibnitz rule that states:
( x, y )

0,0

w
w
dx dy wx, y w0,0
x
y

(10)

Because the point (0, 0) lies on a contour in the solution domain, u at this point is equal to 0 and
v = ln(u) = . Consequently, the right hand side of Equation (10) becomes:
(11)
wx , y w0,0 w x , y e nv w x , y e n ln u w x, y u n w x , y 0
The definite integral on the left side of Equation (10) is calculated at a generic point ( x, y )
identified by the mean of a polygonal curve that starts from (0, 0), passes across ( x , 0 ) and ends at
( x, y ), so that:
y F u
( x , y ) F u
F u 1 1
1 1
ne
dx

ne
d
y

ne 1 1 dy ne 1 1 F u
(12)
0,0 x

0
y
y
in which ( x, y ) represents a point of the solution domain. The right hand side of Equation (12) can be
equated to the right hand side of Equation (11) to obtain:
w x, y ne 1 1 F u

(13)

Because w(x, y) = exp(nv), Equation (13) is rewritten as:


e nv ne 1 F u x , y

(14)

and recalling that v = ln(u), we obtain the expression of u(x,y):

ux, y ne11 Fux, y

1/ n

(15)

Equation (15) contains two Lagrange multipliers 1 and 2, which can be calculated using Equations
(2) and (3). Inserting Equation (4) into Equation (2) and integrating, one obtains:
n
(16)
u max
n e1 1
Combining Equations (15) and (16) we obtain the equation representing the 2D velocity distribution:
(17)
u ( x, y ) u max F 1 / n
Equation (17) is the power law 2D velocity distribution, which depends on umax, n, and the 2D CDF.
The derived equation formally coincides with the equation obtained by Singh [24] for 1D domain,
although in this case F is a function of x and y.
Parameter n can be calculated using the constraint given by Equation (2), resulting in the following
equation [24]:

1
ln u max ln u
n

(18)

where f is given, for the power law equation, by Equation (4). As an alternative, following the
approach proposed by Marini et al. [19], n can be calculated from the definition of average channel
velocity uav:

u av

1
1
udA u max F 1 n dA

A A
A A

(19)

Entropy 2013, 15

1225

where A is the channel cross section. As pointed out in the literature, Equation (18) refers to the mean
of the logarithmic velocity distribution, whereas Equation (19) considers the average logarithmic
velocity in the channel cross section. Unlike , the average channel velocity has a straightforward
physical meaning, and consequently it can be more effective to calculate parameter n.
3. Comparison with Entropy-Based Logarithmic 2D Velocity Distribution

Equation (17) represents an effective way to estimate velocity distribution in a generic 2D domain if
the CDF is properly defined. Starting from the same hypothesis, but using a different constraint
equation than Equation (2), Marini et al. [19] obtained the following equation (logarithmic model):

u max
(20)
ln 1 F e G 1
G
Similar to Equation (17), Equation (19) depends only on umax, an entropic parameter (here called
G), and the 2D CDF. Parameter G can be calculated using the following equation, depending on the
mean of velocity distribution and the maximum velocity [3]:
u x, y

u
u max

eG
1

G
e 1 G

(21)

or considering again the definition of average channel velocity:

1
1 u
(22)
udA max ln 1 F e G 1 dA

A
A
A A G
For both Equations (17) and (20), 1D and 2D domains can be considered, depending on the
assumed CDF. The velocity distributions inferred from the proposed formulation were analyzed for
different configurations and compared with Equation (20) in what follows.
u av

3.1. 1D Velocity Distribution and Maximum Velocity on the Water Level


The case of a wide channel geometry (hence the ratio H/B between the water depth H and the
channel width B tends to zero) is analyzed first. Consequently, only vertical velocity was considered. If
the maximum velocity occurs for y = H, it is well-known [3] that F(u) = y/H. Equation (17) becomes:
u
u max

y

H

1n

(23)

and n can be calculated from the following equation, derived from integration of Equation (19):

u av
u max u av

(24)

If Equation (20) is used with the same F(u), one obtains:

u
u max

y
1
ln1 e G 1
G
H

in which G can be calculated by means of Equation (22).

(25)

Entropy 2013, 15

1226

The following data were considered in the example: H = 1 m; umax = 1 m/s; uav = = 0.8 m/s,
resulting in G = 4.8 from Equation (20) and 1/n = 0.25 from Equation (22). The velocity profiles
obtained from the power law velocity distribution [Equation (17)] and the logarithmic velocity
distribution [Equation (18)] were almost the same (negligible differences), as shown in Figure 1.
Figure 1. Velocity profiles calculated using Equations (17) and (20) for 1D domain and y0 = H.

3.2. 1D Velocity Distribution and Maximum Velocity below the Water Level
In this case, the channel geometry of the previous example was used, except for the point where the
maximum velocity occurs, which is located at a distance y0 = 0.8 m from the bed channel. Also in this
case, the CDF is well-known [5]:
y

y 1 y0
F u
e
y0

(26)

and Equation (17) becomes:


u
u max

y 1 yy

e 0
y0

1n

(27)

and from Equation (19) one obtains 1/n = 0.81. With the same F(u), Equation (20) becomes:
u
u max

y
y 1 y 0
1
G
ln 1 e 1
e
G
y0

(28)

in which G can be calculated, again, using Equation (22), resulting in G = 0.55. The velocity profiles
inferred from the two distributions were almost identical, as shown in Figure 2.
Marini et al. [19] analyzed the same configuration as a special case of the 2D velocity distribution,
thus proposing the following CDF equation:

Entropy 2013, 15

1227
ln 2

ln 2 ln y 0
y

F u 4

2 H

ln 2

y ln 2 ln y 0

2H

(29)

Figure 2. Velocity profiles calculated using Equations (17) and (20) for 1D domain, y0 < H
and F(u) proposed by Chiu [5].

Figure 3 shows the velocity profiles obtained using Equation (27) for F(u), resulting again in
excellent agreement.
Figure 3. Velocity profiles calculated using Equations (17) and (20) for 1D domain, y0 < H
and F(u) proposed by Marini et al. [19].

Entropy 2013, 15

1228

3.3. 2D Velocity Distribution where Maximum Velocity Occurs below Water Level
The 2D domain was analyzed, considering a rectangular channel with symmetrical velocity
distribution with respect to the vertical axis. For this configuration Marini et al. [19] proposed the
following CDF:
2x 2
F u 1
B

H B

ln 2

ln 2 ln y 0
y

2 H

ln 2

y ln 2 ln y 0

2H

(30)

Consequently, Equation (17) becomes:


1/ n

ln 2
ln 2

2 H B

ln 2 ln y 0 H
ln 2 ln y 0 H
y
y
u

2x

1
4

2 H

u max B
2
H

and n can be calculated using Equation (19). With the same F(u), Equation (20) becomes:

(31)

ln 2
ln 2
H B

2x 2
ln 2 ln y 0 H
ln 2 ln y 0 H
y
y
1

(32)

4
ln 1 e 1 1

u max G
B
2
H
2
H

and G can be calculated using Equation (22), in which the average channel velocity can be calculated
from the definition of cross-sectional average velocity for the rectangular channel as:

u av

1
BH

B / 2 H

B / 2 0

udy

(33)

Assuming H = 1 m; B = 1 m; y0 = 0.8 m; umax = 1 m/s; uav = 0.8 m/s, we obtain 1/n = 0.64 from
Equation (19) and G = 1.22 from Equation (30). The velocity profiles at different abscissas were
plotted in Figure 4, resulting in an excellent agreement.
Figure 4. Velocity profiles calculated using Equations (17) and (20) for 2D domain
(rectangular cross section).

Entropy 2013, 15

1229
Figure 4. Cont.

4. Conclusions

Using an entropy-based approach, a power law equation is obtained for the velocity distribution in a
generic 2D domain. The equation formally coincides with that already derived for a 1D domain. For
application, the model only requires the knowledge of average velocity and maximum velocity and the
point where the latter occurs to be known. To assess the reliability of the obtained equation, the
cumulative distribution functions of velocity available in the literature for 1D and 2D domains are used
and velocity distributions in a number of configurations are derived. Results are compared with those
calculated using an entropy-based logarithmic equation available in the literature. For all the analyzed
cases, velocity profiles obtained by the proposed 2D power law velocity distribution show negligible
differences with the logarithmic velocity distribution. Further refinements in the distribution will
investigate the relationship between the power law exponent and the entropic parameter and/or other
physical characteristics of the channel cross section.
References

1.

Singh, V.P. Entropy Theory and Its Applications in Environmental and Water Engineering; John
Wiley: New York, NY, USA, 2013; p. 662.

Entropy 2013, 15
2.

3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.

1230

de Martino, G.; Fontana, N.; Marini, G.; Singh, V.P. Variability and trend in seasonal precipitation
in the continental United States. J. Hydrol. Eng. 2013, 18, doi:10.1061/(ASCE)HE.19435584.0000677.
Chiu, C.L. Entropy and probability concepts in hydraulics. J. Hydraul. Eng. 1987, 113, 583600.
Chiu, C.L. Entropy and 2-D velocity distribution in open channel. J. Hydraul. Eng. 1988, 114,
738756.
Chiu, C.L. Velocity distribution in open channel flow. J. Hydraul. Eng. 1989, 115, 576594.
Chiu, C.L.; Murray, D.W. Variation of velocity distribution along nonuniform open-channel flow.
J. Hydraul. Eng. 1992, 118, 9891001.
Chiu, C.L.; Said, C.A.A. Maximum and mean velocities and entropy in open channel flow.
J. Hydraul. Eng. 1995, 121, 2635.
Chiu, C.L.; Tung, N.C. Maximum velocity and regularities in open channel flow. J. Hydraul. Eng.
2002, 128, 390398.
Chiu, C.L.; Chen, Y.C. An efficient method of discharge estimation based on probability concept.
J. Hydraul. Res. 2003, 41, 589596.
Chiu, C.L.; Hsu, S.M. Probabilistic approach to modeling of velocity distributions in fluid flows.
J. Hydrol. 2006, 316, 2842.
Barbe, D.E.; Cruise, J.F.; Singh, V.P. Solution of three-constraint entropy-based velocity distribution.
J. Hydraul. Eng. 1991, 117, 13891396.
Xia, R. Relation between mean and maximum velocities in a natural channel. J. Hydraul. Eng.
1997, 123, 720723.
Araujo, J.C.; Chaudhary, F.H. Experimental evaluation of 2-D entropy model for open channel
flow. J. Hydraul. Eng. 1998, 124, 10641067.
Kirkgoz, M.S.; Akoz, M.S.; Oner, A.A. Numerical modeling of flow over a chute spillway.
J. Hydraul. Res. 2009, 47, 790797.
Singh, V.P.; Luo, H. Entropy theory for distribution of one-dimensional velocity in open
channels. J. Hydrol. Eng. 2011, 16, 725735.
Luo, H.; Singh, V.P. Entropy theory for two-dimensional velocity distribution. J. Hydrol. Eng.
2011, 16, 303315.
Cui, H.; Singh, V.P. Two-dimensional velocity distribution in open channels using the Tsallis
entropy. J. Hydrol. Eng. 2013, 18, 331339.
Moramarco, T.; Saltalippi, C.; Singh, V.P. Estimation of mean velocity in natural channels based
on Chius velocity distribution equation. J. Hydrol. Eng. 2004, 9, 4250.
Marini, G.; de Martino, G.; Fontana, N.; Fiorentino, M.; Singh, V.P. Entropy approach for 2D
velocity distribution in open channels flow. J. Hydraul. Res. 2011, 49, 784790.
Fontana, N.; Marini, G.; de Paola, F. Experimental assessment of a 2-D entropy-based model for
velocity distribution in open channel flow. Entropy 2013, 15, 988998.
Dingman, S.L. Probability distribution of velocity in natural channel cross sections. Water Res.
1989, 25, 509518.
Shannon, C.E. A mathematical theory of communication. Bell Syst. Tech. J. 1948, 27, 379423,
623656.

Entropy 2013, 15

1231

23. Singh, V.P. Entropy-Based Parameter Estimation in Hydrology; Kluwer Academic Publishers
(now Springer): Dordrecht, The Netherlands, 1998; p. 383.
24. Singh, V.P. Derivation of power law and logarithmic velocity distributions using the Shannon
entropy. J. Hydrol. Eng. 2011, 16, 478483.
25. Jaynes, E.T. Information theory and statistical mechanics, I. Phys. Rev. 1957, 106, 620630.
26. Jaynes, E.T. Information theory and statistical mechanics, II. Phys. Rev. 1957, 108, 171190.
27. Jaynes, E.T. On the rationale of maximum entropy methods. Proc. IEEE 1982, 70, 939952.
2013 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/3.0/).

Вам также может понравиться